首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
Transition metal (M = Mn(II), Co(II), Ni(II) and Cu(II)) complexes with octahydro-Schiff base (H4-N4O4) = 2,7,13,18-tetramethyl-3,6,14,17-tetraazatricyclo-[17.3.1.1]-tetracosa-1(23),2,6,8(24),9,11,13,17,19,21-decaene-9,11,20,22-tetraol; H4([H]8-N4O4) = 2,7,13,,18-tetramethyl-3,6,14,17-tetraazatricyclo-[17.13.1.1.]-tetracosa-1(23),8(24),9,11,19,21-hexane-9,11,20,22-tetraol) have been encapsulated in nanopores of zeolite-Y; [M([H]8-N4O4)]@NaY; with Flexible Ligand Method (FLM) for the first time. The new Host-Guest Nanocomposite Materials (HGNM) was characterized by several techniques: chemical analysis, spectroscopic methods (DRS, FT-IR and UV/Vis), BET technique, conductometric and magnetic measurements. The catalytic activities for oxidation of cyclohexane with HGNM complexes are reported. Zeolite encapsulated octahydro-Schiff base copper(II) complex; [Cu([H]8-N4O4)]@NaY; was found to be more active than the corresponding cobalt(II), manganese(II) and nickel(II) complexes for cyclohexane oxidation. The catalytic properties of the complexes are influenced by their geometry and by the steric environment of the active sites. HGNM are stable enough to be reused and are suitable to be utilized as partial oxidation catalysts. The encapsulated catalysts systems; [M([H]8-N4O4)]@NaY; were more active than the corresponding neat complexes; [M([H]8-N4O4))].  相似文献   

2.
Formation constants of ternary complexes MAL, where M = Cu(II) or Ni(II). A = 2.2′bipyridyl. 1, 10-phenanthroline, and L = 3.4-dihydroxyphenylalanine (dopa), tyrosine, or phenylalanine have been determined by using the computer program SCOGS. It is observed that dopa coordinates with Cu(II)-A and Ni(II)-A through the aminocarboxylate and only over the pH range 3–8, though the ligand coordinates with free Cu(II) ion from the amino carboxylate end in the lower pH range (pH 2–4) and from the catechol end at the higher pH range (pH > 5). The visible spectrum of Cu-A-dopa is similar to that of Cu-A-phenylalanine or Cu-A-tyrosine over the entire pH range, confirming amino carboxylate coordination. Δ log K (KMAL - logKML) is found to be positive in all the six Cu(II) complexes. whereas it is negative in Ni(II) complexes. Release in the ternary complexes of the repulsion between the Cu(II) dπ electron and electrons delocalized over the phenyl ring has been proposed as a probable reason for the positive Δ log K.  相似文献   

3.
The presence of the bicarbonate/carbon dioxide pair is known to accelerate the transition metal ion-catalysed oxidation of various biotargets. It has been shown that stable Cu(II) complexes formed with imine ligands that allow redox cycling between Cu(I) and Cu(II) display diverse apoptotic effects on cell cultures. It is also reported that Cu(II)-tetraglycine can form a stable Cu(III) complex. In the present study, radical generation from H2O2 and H2O2/HCO3 in the presence of these two different classes of Cu(II) complexes was evaluated by monitoring the oxidation of dihydrorhodamine 123 and NADH and by the quantitative determination of thiobarbituric acid reactive substances (TBARs method). Cu(II)-imine complexes produced low levels of reactive species whereas Cu(II)-Gly-derived complexes, as well as the free Cu(II) ion, produced oxygen-derived radicals in significantly larger amounts. The effects of these two classes of complexes on mammalian tumour cell viability were equally distinct, in that Cu(II)-imine complexes caused apoptosis, entered in cell and remained almost unaffected in high levels whilst, at the same concentrations, Cu(II)-Gly peptide complexes and Cu(II) sulphate stimulated cell proliferation, with the cell managing copper efficiently. Taken together, these results highlight the different biological effects of Cu(II) complexes, some of which have been recently studied as anti-tumour drugs and radical system generators, and also update the effects of reactive oxygen species generation on cell cycle control.  相似文献   

4.
Several features of the catalytic oxidation of cysteine by ceruloplasmin and nonenzymic Cu(II) at pH 7 have been compared. The oxidation of cysteine by ceruloplasmin has several properties in common with the Cu(II) catalyzed oxidation of cysteine: pH maxima, thiol specificity, lack of inhibition by anions, and high sensitivity to inhibition by copper complexing reagents. These two catalysts differed in their molecular activity, in their ability to oxidize penicillamine and thioglycolate, and in that H2O2 was produced as a primary product only during Cu(II) oxidation. The oxidation of cysteine by ceruloplasmin was compared also with the ceruloplasmin catalyzed oxidation of o-dianisidine, a classical pH 5.5 substrate. The mechanism of the oxidation of cysteine by ceruloplasmin at pH 7 differed from that of o-dianisidine oxidation because the latter substrate was inhibited by anions but not by copper complexing agents. Spectral and other data suggest that during the ceruloplasmin reaction with cysteine there is a one electron transfer from cysteine to ceruloplasmin resulting in the specific reduction of type lb Cu(II).  相似文献   

5.
A number of di-Cu(II) complexes of the new tetraimine macrocyclic ligand derived from the Schiff base [2 + 2] condensation of 2,5-diformylfuran with 3-oxa-pentane-1,5-diamine have been prepared by methods which employ the heavier alkaline earth metal ions as templates followed by transmetallation. The complexes have been characterised by spectroscopic and other physical methods. Several of the di-Cu(I) complexes react reversibly with CO in solution and irreversibly with O2 in a 4:1 Cu:O2 stoicheiometry. Depending on conditions the oxidation product may be a dinuclear Cu(II) complex of the macrocycle or a mononuclear Cu(II) complex of a new ring-opened ligand. The single crystal X-ray structure of the latter complex has been determined.[CuL](BPh4)2 is monoclinic, space group C2/c with a=20.12(1), b=14.48(1), c=22.37(2) Å, β=110.1(1)°, Z=4. 1389 Independent reflections above background were measured on a diffractometer and the structure refined to R=0.108. The cation has imposed C2 symmetry. The copper atom is bonded to four nitrogen atoms in the ‘outer’ compartment of the ligand with unique CuN distances of 2.050(17) and 1.977(17) Å. The geometry of the copper atom is intermediate between square planar and tetrahedral with an angle of 39.7° between two CuN2 planes. Molecular mechanics calculations show that this distortion is due to steric effects.  相似文献   

6.
Chiral N,O pyridine alcohols HL1-HL6 were used to form complexes with copper(II) ions. Ligands HL1 and HL2 formed complexes with copper(II) ions when Cu(OAc)2 and HL were refluxed in methanol/ethanol mixture. Ligand HL3 formed a complex with copper(II) when deprotonated with NaH and stirred in a Cu(II) acetate THF solution. Ligands HL4-HL6 did not form complexes with copper(II) under similar conditions. Two complexes, [Cu(L1)2] and [Cu(L2)2], were isolated as single crystals and characterized by X-ray crystallography. These complexes showed low catalytic activities in asymmetric reactions. However, they became active when reacted with triflic acid. Copper complexes, [Cu(L)] or [Cu(L)]+, formed in situ by reacting ligands HL with copper(I) or (II) ions, respectively, were also found to be active copper catalysts for asymmetric cyclopropanation of styrene with ethyl diazoacetate and allylic oxidation of cyclohexene with t-butylperoxybenzoate. Enantioselectivities up to 56% and 38% were obtained in asymmetric cyclopropanation of styrene and asymmetric allylic oxidation of cyclohexene, respectively.  相似文献   

7.
Copper(II) complexes of a new bis benzimidazole diamide ligand N-picolyl-N,N′-bis(2-methylbenzimidazolyl)hexanediamide [Pic-GBHA = L2] have been synthesized and characterized. One of the compound [Cu(L2)(NO3)2] has been structurally characterized. The copper atom is bound to two benzimidazolyl nitrogen atoms, two amide carbonyl oxygen atoms and a bidentate nitrate ion, resulting in a distorted octahedral geometry. EPR spectra obtained at low temperature indicate a tetragonal geometry in the solution state. Complexes display a quasi-reversible redox wave due to the Cu(II)/Cu(I) reduction process having fairly cathodic E1/2. These Cu(II) complexes were utilized to carry out oxidation of ditertbutylcatechol (DTBC) in methanol using molecular oxygen as the oxidant in. Low temperature EPR study of the oxidation reaction implicates the formation of an active copper species with fairly low A value. The presence of picolyl groups on the ligand also serve as a proton sponge giving 2-3 times higher rates of reaction in comparison to the non-picolylated ligand, implying a role of free basic groups in the pH control of enzymatic oxidation of catechols by catechol oxidase and tyrosinase.  相似文献   

8.
9.
When lysozyme was treated with Cu(II) and H2O2 at pH 7.4, the protein underwent polymerization as well as changes in its fluorescent characteristics. Upon prolonged incubation, most of the protein aggregates were degraded into smaller peptides. Amino acid analysis indicated that the basic amino acid residues were most susceptible to the oxidation. Tryptophan residues were converted to N-formylkynurenine and kynurenine, and lysine residues were deaminated to form α-aminoadipic acid δ-semi- aldehyde. During Cu(II)H2O2 treatment, the formation of carbonyl groups was accompanied by the loss of free amino groups in the protein. Succinylation of free amino groups protected lysine residues from oxidation by Cu(II)H2O2, but failed to prevent polymerization. The studies with the modified lysozyme suggest that Cu(II)H2O2 can oxidize various amino acid residues in addition to lysine to generate different types of carbonyl compounds and these carbonyl compounds may be responsible for the formation of crosslinks in the polymerization process.  相似文献   

10.
11.
Formation (affinity) constants for 1:1 complexes of N-(2-acetamido)iminodiacetic acid (ADAH2) with Ca(II), Mg(II), Mn(II), Zn(II), Co(II), Ni(II), and Cu(II) have been determined. Probable structures of the various metal chelates existing in solution are discussed. Values for the deprotonation of the amide group in [Cu(ADA)] and subsequent hydroxo complex formation are also reported. The use of ADA as a buffer is considered in terms of metal buffers complexes which can be formed at physiological pH, i.e., at pH 7.0 there is essentially no free metal ion in 1:1 M2+ to ADA solutions.  相似文献   

12.
Electron spin resonance spectra of the first Cu(II) complexes of human serum albumin, dog serum albumin, l-aspartyl-l-histidine N-methylamide and glycyl-glycyl-l-histidine N-methylamide have been studied using isotopically pure 65Cu in its chloride form. At 77° K, the esr spectra of Cu(II) complex of human serum albumin exhibited only one form of esr signal between pH 6.5 and 11. No intermediate forms were detected. The presence of an equally spaced nine-line superhyperfine structure with spacing ~15 G indicated considerable covalent bonding between Cu(II) and four nitrogen atoms derived from the protein. The esr spectrum form of Cu(II) bound to human serum albumin detected at neutral pH would be consistent with the participation of four nitrogens from the α-NH2 group, two peptide groups, and the imidazole group of a histidine residue. In contrast, the esr spectra of Cu(II)-dog serum albumin complex showed a transition from a low pH form to a high pH form as the pH was increased to 9.5. These spectral changes were found to be reversible upon lowering the pH. Ligand superhyperfine splittings in the low pH form of the esr signal of Cu(II)-dog albumin were not resolved. The distinct pH dependence of the esr signals observed in human and dog serum albumin complexes could be correlated to their respective optical spectra changes as a function of pH. At room temperature and in the pH range between 6 and 11, the esr spectra of Cu(II) complexes of l-aspartyl-l-alanyl-l-histidine N-methylamide and glycyl-glycyl-l-histidine N-methylamide exhibited a well-resolved nine-line superhyperfine structure indicating metal coordination with four equivalent nitrogen atoms of peptide.  相似文献   

13.
The electronic effects of the fluorine atoms in hfacac (hexafluoroacetylacetonato) compared with acac (acetylacetonato) in Cu(II) complexes are used to control the molecular and supramolecular structure of Cu(II) compounds. While bis(acac)Cu(II) (acac = acetylacetonato) is known to be able to have a fifth-position coordination, bis(hfacac)Cu(II), (hfacac = hexafluoroacetylacetonato) may have two extra ligands. This, together with the reliable “supramolecular reagent” isonicotinamide, as the additional ligand, are used to go from a zero-dimension structure, with Cu-acac, to an extended supramolecular two-dimension network, with Cu-hfacac. The molecular and crystal structure of bis(acetylacetonato-O,O′)-(isonicotinamide-N) copper(II), 1, and bis(hexafluoroacetylacetonato-O,O′)-trans-bis(isonicotinamide-N) copper(II), 2, are reported.  相似文献   

14.
Complexes formed between Mn(II) ion and acetohydroxamic acid (HAha), benzohydroxamic acid (HBha), N-methyl-acetohydroxamic acid (HMeAha), DFB model dihydroxamic acids (H2(3,4-DIHA), H2(3,3-DIHA), H2(2,5-DIHA), H2(2,5-H,H-DIHA), H2(2,4-DIHA), H2(2,3-DIHA)) and two trihydroxamate based natural siderophores, desferrioxamine B (H4DFB) and desferricoprogen (H3DFC) have been investigated under anaerobic condition (and some of them also under aerobic condition). The pH-potentiometric results showed the formation of well-defined complexes with moderate stability. Monohydroxamic acids not, but all of the dihydroxamic acids and trihydroxamic acids were able to hinder the hydrolysis of the metal ion up to pH ca. 11. Maximum three hydroxamates were found to coordinate to the Mn(II) ion, but presence of water molecule in the inner-sphere was also indicated by the corresponding relaxivity values even in the tris-chelated complexes. Moreover, prototropic exchange processes were found to increase the relaxation rate of the solvent water proton over the value of [Mnaqua]2+ in the protonated Mn(II)-siderophore complexes at physiological pH. The much higher stability of Mn(III)-hydroxamate (especially tris-chelated) complexes compared to the corresponding Mn(II)-containing species results in a significantly decreased formal potential compared to the Mn(III)aqua/Mn(II)aqua system. As a result, air oxygen becomes an oxidizing agent for these manganese(II)-hydroxamate complexes above pH 7.5. The oxidation processes, followed by UV-Vis spectrophotometry, were found to be stoichiometric only in the case of the tris-chelated complexes of siderophores, which predominate above pH 9. ESI-MS provided support about the stoichiometry and cyclic-voltammetry was used to determine the stability constants for the tris-chelated complexes, [Mn(HDFB)]+ and [MnDFC].  相似文献   

15.
Interactions of inosine derivatives with copper(II) were studied in the pH range 1.4–13 in 50% H2O-50% DMSO solution. The distinct pH dependence of the optical spectra observed in copper(II)-inosine complexes are correlated to their respective EPR changes as a function of pH. It was concluded that a simple 1:1 complex of copper(II)-inosine is formed in the pH range 1.4–5.0 and bis complexes are present in the pH 5.0–6.2 region solutions of inosine and Cu(II). From pH 6.2 to 7.8 a diamagnetic, hydroxybridged complex dominates. At pH 7.8–9.2 an insoluble, oxybridged species is formed in addition to the soluble paramagnetic Cu(NI)4 complex. Starting from pH 9.1 the N-polymeric complex is formed which is stable up to pH 12.5, and above pH 12.5 the only species is the Cu(ribose)2 complex.  相似文献   

16.
Ascorbic acid (vitamin C) induced hydrogen peroxide (H2O2) formation was measured in household drinking water and metal supplemented Milli-Q water by using the FOX assay. Here we show that ascorbic acid readily induces H2O2 formation in Cu(II) supplemented Milli-Q water and poorly buffered household drinking water. In contrast to Cu(II), iron was not capable to support ascorbic acid induced H2O2 formation during acidic conditions (pH: 3.5–5). In 12 out of the 48 drinking water samples incubated with 2 mM ascorbic acid, the H2O2 concentration exceeded 400 μM. However, when trace amounts of Fe(III) (0.2 mg/l) was present during incubation, the ascorbic acid/Cu(II)-induced H2O2 accumulation was totally blocked. Of the other common divalent or trivalent metal ions tested, that are normally present in drinking water (calcium, magnesium, zinc, cobalt, manganese or aluminum), only calcium and magnesium displayed a modest inhibitory activity on the ascorbic acid/Cu(II)-induced H2O2 formation. Oxalic acid, one of the degradation products from ascorbic acid, was confirmed to actively participate in the iron induced degradation of H2O2. Ascorbic acid/Cu(II)-induced H2O2 formation during acidic conditions, as demonstrated here in poorly buffered drinking water, could be of importance in host defense against bacterial infections. In addition, our findings might explain the mechanism for the protective effect of iron against vitamin C induced cell toxicity.  相似文献   

17.
Although quinones represent a class of organic compounds that may exert toxic effects both in vitro and in vivo, the molecular mechanisms involved in quinone species toxicity are still largely unknown, especially in the presence of transition metals, which may both induce the transformation of the various quinone species and result in generation of harmful reactive oxygen species. In this study, the oxidation of 1,4-naphthohydroquinone (NH2Q) in the absence and presence of nanomolar concentrations of Cu(II) in 10 mM NaCl solution over a pH range of 6.5–7.5 has been investigated, with detailed kinetic models developed to describe the predominant mechanisms operative in these systems. In the absence of copper, the apparent oxidation rate of NH2Q increased with increasing pH and initial NH2Q concentration, with concomitant oxygen consumption and peroxide generation. The doubly dissociated species, NQ2−, has been shown to be the reactive species with regard to the one-electron oxidation by O2 and comproportionation with the quinone species, both generating the semiquinone radical (NSQ). The oxidation of NSQ by O2 is shown to be the most important pathway for superoxide (O2) generation with a high intrinsic rate constant of 1.0×108 M−1 s−1. Both NSQ and O2 served as chain-propagating species in the autoxidation of NH2Q. Cu(II) is capable of catalyzing the oxidation of NH2Q in the presence of O2 with the oxidation also accelerated by increasing the pH. Both the uncharged (NH2Q0) and the mono-anionic (NHQ) species were found to be the kinetically active forms, reducing Cu(II) with an intrinsic rate constant of 4.0×104 and 1.2×107 M−1 s−1, respectively. The presence of O2 facilitated the catalytic role of Cu(II) by rapidly regenerating Cu(II) via continuous oxidation of Cu(I) and also by efficient removal of NSQ resulting in the generation of O2. The half-cell reduction potentials of various redox couples at neutral pH indicated good agreement between thermodynamic and kinetic considerations for various key reactions involved, further validating the proposed mechanisms involved in both the autoxidation and the copper-catalyzed oxidation of NH2Q in circumneutral pH solutions.  相似文献   

18.
We present evidence that the oxidation of Mn(II) in a zone above the O2/H2S interface in the water column of Saanich Inlet, British Columbia, Canada, is microbially catalyzed. We measured the uptake of 54Mn(II) in water samples under in situ conditions of pH and temperature and in the presence and absence of oxygen. Experiments in the absence of oxygen provided a measure of the exchange of the tracer between the dissolved and solid pools of Mn(II); we interpret the difference between experiments in the presence and absence of oxygen to be a measure of Mn(II) oxidation. Using this method we examined the effect of oxygen tension, Mn(II) concentration, and temperature on the initial in situ Mn(II) oxidation rate (V0). Mn(II) oxidation was almost twice as fast under conditions of 67% air saturation (V0=5.5 nM h−1) as with the in situ concentration of 15 μM (5% air saturation; V0=3.1 nM h−1). Additions of ca. 18 μM Mn(II) completely inhibited all Mn(II) oxidation at three different depths in the oxidizing zone, and there was a temperature optimum for Mn(II) oxidation of around 20°C. These results are consistent with biologically mediated Mn(II) oxidation and indicate that the rate is limited by both oxygen and the concentration of microbial binding sites in this environment.  相似文献   

19.
The Schiff base ligands 2-(2,6-diisopropylphenyliminomethyl)phenol H(L1), 5-diethylamino-2-(2,6-diisopropylphenyliminomethyl)phenol H(L2), 2,4-di-tert-butyl-6-(2,6-diisopropylphenyliminomethyl)phenol H(L3), 3-(2,6-diisopropylphenyliminomethyl)naphthalen-2-ol H(L4) and 4-(2,6-diisopropylphenyliminomethyl)-5-hydroxymethyl-2-methylpyridin-3-ol H(L5) have been synthesized by the condensation, respectively, of salicylaldehyde, 4-(diethylamino)salicylaldehyde, 3,5-di-tert-butylsalicylaldehyde, 2-hydroxy-1-napthaldehyde and pyridoxal with 2,6-diisopropylaniline. The copper(II) bis-ligand complexes [Cu(L1)2] 1, [Cu(L2)2] 2, [Cu(L3)2] 3, [Cu(L4)2] 4 and [Cu(L5)2] · CH3OH 5 of these ligands have been isolated and characterized. The X-ray crystal structures of two of the complexes [Cu(L1)2] 1 and [Cu(L5)2] · CH3OH 5 have been successfully determined, and the centrosymmetric complexes possess a CuN2O2 chromophore with square planar coordination geometry. The frozen solution EPR spectra of the complexes reveal a square-based CuN2O2 chromophore, and the values of g and g/A index reveal enhanced electron delocalization by incorporating the strongly electron-releasing -NEt2 group (2) and fusing a benzene ring on sal-ring (4). The Cu(II)/Cu(I) redox potentials of the Cu(II) complexes reveal that the incorporation of electron-releasing -NEt2 group and fusion of a benzene ring lead to enhanced stabilization of Cu(II) oxidation state supporting the EPR spectral results. The hydrogen bonding interactions between the two molecules present in the unit cell of 5a generate an interesting two-dimensional hydrogen-bonded network topology.  相似文献   

20.
The interaction of the cupric ion with phytic acid as a function of pH has been studied by potentiometric and thermal titration and by the determination of ligand binding. As has been found for the reaction of zinc and calcium cations with phytate, the presence of the Cu(II) ion results in a displacement of the titration curves to more acid values. Evaluation of the parameters that describe such changes in ionization behavior by curve-fit analysis showed that as the Cu(II):phytate mol ratio was increased from one to eight, the pK' values of the ionizable group sets of phytic acid (ranging from 1.59 to 9.79) were consolidated into just two sets with curve-fit (CP) values ca. 1.5 and 3.7. Marked pH hysteresis effects are seen in such systems because of the pronounced acid strength of the Cu(II):aqua ion and the Cu(II) ligand aqua ion complex. The combined heat of binding and precipitation (plus solvation changes, etc.) of Cu(II) to phytate is endothermic (21.8–22.2 kcal mol−1). This is similar in magnitude to that reported for the binding of either Zn(II) or Ca(II) to phytate. In the titration of Cu(NO3)2 with KOH, presumably to form Cu(OH)2, ΔH° was exothermic (−12.5 kcal mol−1). From measurements of free Cu(II) cation concentration in the presence of phytate the binding reaction was found to be stoichiometric with 6 mols Cu(II) bound at pH 6. Binding occurs within the pH range 2–6. An apparent necessary requirement for binding is the availability of the oxo dianion structure formed from the second dissociation step of a phosphoryl group. Curve-fit analysis of the binding data as a function of pH showed that a group or group set with CP value ca. 4 governs the binding reaction(s) at all mol ratios of Cu(II) to phytate examined. It is suggested that the binding of cupric ions to phytate may occur to the equatorial rather than the axial configuration as suggested for Ca(II) binding. A space-filling molecular model to illustrate this has been constructed. Soluble Cu(II):phytate complexes are formed within the pH range from 2 to ca. 3.4. This is supported by the results of difference absorption spectrometry.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号